Skip to main content
Cold Spring Harbor Perspectives in Biology logoLink to Cold Spring Harbor Perspectives in Biology
. 2012 Apr;4(4):a005686. doi: 10.1101/cshperspect.a005686

Calcium Signaling in Dendritic Spines

Michael J Higley 1, Bernardo L Sabatini 2
PMCID: PMC3312680  PMID: 22338091

Abstract

Calcium (Ca2+) is a ubiquitous signaling molecule that accumulates in the cytoplasm in response to diverse classes of stimuli and, in turn, regulates many aspects of cell function. In neurons, Ca2+ influx in response to action potentials or synaptic stimulation triggers neurotransmitter release, modulates ion channels, induces synaptic plasticity, and activates transcription. In this article, we discuss the factors that regulate Ca2+ signaling in mammalian neurons with a particular focus on Ca2+ signaling within dendritic spines. This includes consideration of the routes of entry and exit of Ca2+, the cellular mechanisms that establish the temporal and spatial profile of Ca2+ signaling, and the biophysical criteria that determine which downstream signals are activated when Ca2+ accumulates in a spine. Furthermore, we also briefly discuss the technical advances that made possible the quantitative study of Ca2+ signaling in dendritic spines.


Synaptically evoked Ca2+ transients in dendritic spines are shaped by a complex interplay among Ca2+ sources, pumps, and binding proteins, as well as by the biophysics of Ca2+ diffusion and spine morphology.


For many neurons in the mammalian brain, the postsynaptic terminal of an excitatory synapse is found in a specialized structure protruding from the dendritic shaft, known as a dendritic spine (Fig. 1). The structure of the classic mushroom-shaped dendritic spine, in which a bulbous head (diameter ∼0.5 µm) is separated from the parent dendritic shaft by a thin neck (length ∼0.5 µm, diameter ∼0.1 μm), suggests that it creates an isolated signaling compartment in which the machinery necessary to read out and regulate the activity of one synapse can operate independently of that associated with a neighboring synapse. Indeed, many studies have shown that the spine neck provides a significant barrier to diffusion that allows compartmentalization of biochemical and electrical signals in the spine head (Yuste and Denk 1995; Svoboda et al. 1996; Sabatini et al. 2002; Grunditz et al. 2008; Bloodgood et al. 2009). The generalizability of these conclusions to non-mushroom spines (see Harris and Weinberg 2012), such as those with no discernible neck (stubby spines) or those with a long neck and small head (thin spines), is still unclear.

Figure 1.

Figure 1.

Sources and regulation of Ca2+ in dendritic spines. Synaptic stimulation produces postsynaptic depolarization via currents (black arrows) through AMPA- and NMDA-type glutamate receptors as well as various voltage-gated ion channels. Synaptic Ca2+ influx (blue arrows) is mediated by NMDARs and voltage-gated Ca2+ channels (VGCCs) and is possibly augmented by release from intracellular stores such as smooth endoplasmic reticulum (SER). The membrane potentials within the spine (Vspine) and dendritic shaft (Vdend) are electrically coupled by the spine neck. Similarly, the Ca2+ concentrations in the spine ([Ca]spine and dendritic shaft [Ca]dend) are coupled by restricted diffusion across the neck. The presence of Ca-activated (SK) potassium channels in the spine head provides a negative-feedback loop regulating synaptic depolarization. Sodium (INa) and potassium (IK) channels, as well as SER, are shown only in the dendritic shaft for simplicity but may also be present in the spine head. Note that for clarity the spine neck is not drawn to scale.

The best-studied, synaptically evoked biochemical signal that accumulates in active spines is intracellular calcium (Ca2+) (for review, see Sabatini et al. 2001; Bloodgood and Sabatini 2007a; Higley and Sabatini 2008). In pyramidal neurons of the neocortex and hippocampus, synaptic activation of NMDA-type glutamate receptors (NMDARs) located in the postsynaptic density (PSD) leads to influx of Ca2+, which accumulates in the head of the spine associated with the active synapse (Fig. 1). Additional Ca2+ enters through voltage-gated Ca2+ channels (VGCCs) or may be released from intracellular Ca2+ stores such as mitochondria and endoplasmic reticulum. Current carried by Ca2+ ions contributes to dendritic electrical signaling, producing postsynaptic depolarization. Additionally, Ca2+ ions in the spine head activate a wide variety of Ca-sensitive proteins—including calmodulin, Ca2+/calmodulin-dependent kinase type II (CaMKII), small conductance Ca2+-activated potassium channels, calcineurin, and calpain—that regulate many aspects of neuron and synapse function. Thus, the Ca2+ that accumulates in the spine is a central signaling molecule that regulates many aspects of synapse and cell function.

TOOLS TO STUDY Ca2+ SIGNALING IN DENDRITIC SPINES

The explosion in our quantitative understanding of Ca2+ signaling in dendritic spines has been made possible by three fundamental technical advances that are briefly discussed here (for review, see Yasuda et al. 2004). These are the development of bright, fast, and high-dynamic-range Ca2+-sensitive fluorophores; the development and dissemination of two-photon laser-scanning microscopy; and the development of two-photon released caged neurotransmitters that allow direct stimulation of visualized spines (Tsien 1980; Minta et al. 1989; Denk et al. 1990; Matsuzaki et al. 2001; Carter and Sabatini 2004; Yasuda et al. 2004).

Ca2+ accumulation within cells and dendritic spines can be visualized through the use of Ca2+-sensitive fluorescent molecules that are commonly referred to as “Ca2+ indicators.” The current generation of these molecules typically consists of a Ca2+-binding molecule, such as the common Ca2+ buffer 1,2-bis(o-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid (BAPTA), that has been attached to a fluorophore such as fluorescein (Tsien 1980; Minta et al. 1989). Ca2+ binding to the buffer moiety triggers a change in the electronic properties of the fluorophore that alters its fluorescence, typically by altering its quantum yield or absorption cross section (Kao 1994; Wokosin et al. 2004). In this case, intracellular Ca2+ accumulation alters the emission of green fluorescence, which can be detected by CCD cameras or photomultiplier tubes (Fig. 2). Ca2+ indicators are now available that differ widely in their properties. For example, although indicators most often used for the study of dendritic spines fluoresce green light (e.g., Fura-2, Oregon Green-BAPTA, Fluo-4), some indicators emit blue or red light (e.g., X-Rhod). In addition, the affinities of available indicators for Ca2+ vary over several orders of magnitude, from ∼10−7 to 10−3 m. Last, although the ease of use, fast Ca2+ binding, and brightness of synthetic Ca2+ indicators have made them the gold standard, the progress in developing genetically encoded Ca2+ indicators may soon yield designer proteins suitable for the study of synaptic Ca2+ signaling (Miyawaki et al. 1997; Nagai et al. 2001; Mank et al. 2006; Tian et al. 2009). By careful consideration of the properties of the Ca2+ transient under study, it is typically possible to choose a Ca2+ indicator and the concentration at which it is used such that the magnitude of fluorescence change is directly and nearly linearly related to the change in intracellular Ca2+ concentration.

Figure 2.

Figure 2.

Ca2+-sensitive fluorophores and two-photon excitation. (A) The structure of the Ca2+-sensitive fluorophore Fluo-4 is shown (lower left). This molecule consists of aromatic moiety, based on fluorescein, connected to a negatively charged Ca2+-binding molecule, based on BAPTA. When Ca2+ binds to the buffer via coordination of the negative charges (upper molecule), the quantum yield, and thus fluorescence, of Fluo-4 increases dramatically. In the Ca2+-bound state, Fluo-4 can be excited by a single blue photon (not shown) or via near-simultaneous absorption of two low-energy red photons (i.e., two-photon stimulation). In either case, the emission consists of a single green photon. (B) Simple schematic of a two-photon scanning microscope. Infrared (∼800 nm) light (red line) is generated by a high-energy titanium sapphire laser and directed into the microscope objective via a light path consisting of a scanning galvanometer mirror system that scans the laser in two dimensions and a telescope. The mirrors are used to scan the excitation beam through the sample. Visible fluorescence (green line) produced at the sample is collected and directed into a photomultiplier tube for image formation. (C) Schematic illustrating two-photon imaging of Ca2+ transients evoked by simultaneous two-photon glutamate uncaging. A neuron is filled with a Ca2+ indicator, such as Fluo-4, via the recording pipette and bathed in the inactive but photolabile compound MNI-glutamate. One laser (840 nm) is scanned across the tissue to image fluorescence. A second laser (725 nm) is directed next to the spine head, where a brief pulse breaks the covalent bond between glutamate and the blocking moiety. The released glutamate molecules bind synaptic NMDARs and lead to Ca2+ influx. (D) Example of uncaging-evoked synaptic Ca2+ transient in a dendritic spine. The neuron has been filled with the Ca2+-insensitive red fluorophore Alexa Fluor-594 to reveal spine morphology and the Ca2+-sensitive green fluorophore Fluo-4. Two-photon glutamate uncaging at the indicated location (arrowhead) evokes a brief Ca2+ transient (ΔCa) and corresponding excitatory postsynaptic potential (EPSP).

However, despite the relative ease of use of Ca2+ indicators, great care must be taken in the analysis of Ca2+-dependent changes in the fluorescence emitted by a Ca2+ indicator. Because Ca2+ indicators bind Ca2+ and are often present at high concentrations, from a biochemical signaling point of view, they are functionally equivalent to a Ca2+ buffer such as BATPA or EGTA. Hence, their presence fundamentally perturbs the Ca2+ transient to be analyzed, typically greatly reducing its amplitude, increasing its duration, and accelerating its spatial spread (Box 1). Furthermore, the blunting effect they have on the amplitude of evoked Ca2+ transients typically prevents or dampens the activation of downstream Ca2+-dependent processes. Thus, it is generally difficult to study both a Ca2+-dependent process and the triggering Ca2+ transient at the same time. Several studies have discussed quantitative descriptions of the perturbing properties of Ca2+ buffers and practical considerations in selecting a Ca2+ indicator (Neher and Augustine 1992; Tank et al. 1995; Neher 1998; Sabatini et al. 2001; Yasuda et al. 2004; Higley and Sabatini 2008).

BOX 1. Ca2+ BUFFERING BY ENDOGENOUS AND EXOGENOUS MOLECULES.

Intracellular calcium (Ca) reversibly binds to endogenous molecules such as the Ca2+-binding proteins calbindin, parvalbumin, and calmodulin. In addition, Ca2+ binds to experimentally introduced molecules including buffers such as EGTA and BAPTA, synthetic Ca2+ indicators such as Fura-2 and Fluo-4, and the genetically encoded Ca2+ indicators GCAMP and TN-XL. Although exogenously applied Ca2+ buffers have been used extensively as tools to study Ca2+ function, particularly when used as fluorescent indicators of intracellular Ca2+ concentration, all buffers (both endogenous and exogenous) alter the dynamics of Ca2+ signaling. Understanding the nature of this perturbation is critical to the proper interpretation of data derived with these tools.

The effects of buffering on the dynamics of intracellular Ca2+ concentration have been quantitatively described (Neher and Augustine 1992; Zhou and Neher 1993). The equilibrium between free Ca2+, molecules of unbound buffer (B), and Ca2+-bound buffer (BCa) is described by the laws of mass action as:

[Ca]+[B][BCa]

with equilibrium dissociation constant:

KD=[Ca][B][BCa]

The specific relationship between a change in the concentration of free Ca2+ concentration and the corresponding change in the concentration of Ca2+-bound buffer is described quantitatively by the buffer capacity, κB, defined as the incremental change in BCa2+ for an incremental change in free Ca2+:

κB=δ[BCa]δ[Ca]

This relationship defines the slope of the familiar saturating curve relating the concentrations of free Ca2+ and Ca2+-bound buffer (Fig. 3). The nonlinearity of the curve indicates that κB is not constant but decreases as Ca2+ increases. Intuitively, at higher Ca2+ concentrations, a greater fraction of the total buffer is bound, and less is available to sequester further increases in Ca2+. The buffer capacity κB is directly proportional to Ca2+ affinity (or inversely proportional to KD) and the concentration of the buffer. Thus, high-affinity buffers strongly sequester free Ca2+, but are also saturated more readily (Higley and Sabatini 2008).

Figure 3.

Figure 3.

Curve showing the relationship between the concentrations of free Ca2+ and Ca2+-bound buffer.

In response to a brief increase in total cytosolic Ca2+ (such as occurs during a synaptic event or action potential), the presence of exogenous buffer produces several perturbations in the dynamics of free Ca2+ concentration compared with what would occur in the absence of added buffers. The degree of perturbation is directly proportional to the fractional buffer capacity that comes from exogenous buffers.

First, the added buffer temporarily sequesters Ca2+ ions, preventing their interaction with other intracellular molecules. Quantitatively, the change in free Ca2+ is equal to the change in total Ca2+ scaled by the sum of the capacities of each buffer in the cell. Or, considering simply one exogenous buffer with capacity κE and total native buffer capacity κN:

Δ[Ca]=ΔCaT1+κE+κN

Thus, relative to the situation without added buffers (i.e., κE = 0), the amplitudes of evoked Ca2+ transients are reduced in proportion to the relative increases in buffer capacity. Because 100 µm of EGTA or BAPTA provides a buffer capacity of ∼250 and spines have native buffer capacities of ∼20 (Sabatini et al. 2002), the exogenous buffers can very quickly dominate and reduce the amplitude of free Ca2+ transients by more than a factor of 10. The functional consequences of this reduction are well known to investigators of Ca2+-dependent long-term synaptic plasticity, as the experimental introduction of EGTA or BAPTA can reduce or prevent plasticity induction as well as many other Ca2+-dependent processes.

Second, somewhat counter-intuitively, the presence of buffer prolongs the time that Ca2+ is present in the cytosol by slowing the time course of Ca2+ clearance. Free Ca2+ is primarily cleared by extrusion through the cell membrane or by pumping into intracellular storage sites such as mitochondria and endoplasmic reticulum. By preventing Ca2+ ions from binding to pumps, buffers increase the total time necessary to remove Ca2+ from the cytosol. Quantitatively, the time constant (τ) of the exponential time course of decay of free Ca2+ is inversely proportional to the rate of clearance (γ) and directly proportional to the total buffer capacity:

τ=1+κE+κNγ

Thus, the fractional slowing of Ca2+ clearance is proportional to the relative change in buffer capacity caused by the addition of an exogenous buffer. Furthermore, cells with higher native buffer capacity will, given a fixed membrane concentration of Ca2+ pumps, clear Ca2+ more slowly.

Third, the presence of buffer expands the spatial extent of Ca2+ signaling in the cytosol. Intracellular Ca2+ signaling is spatially limited because the spread of Ca2+ ions is limited by binding to endogenous large Ca2+-binding proteins (Soler-Llavina and Sabatini 2006; Schmidt et al. 2007). In contrast, exogenous Ca2+ buffers are typically small (∼500 Da), charged molecules and are highly mobile in the intracellular space. Thus, Ca2+ “catches a ride” while bound on the mobile buffer, bypassing the Ca2+-binding proteins and ultimately unbinding at a more distal location. Quantitatively, the effective diffusion coefficient for Ca2+ (Deff) in the presence of buffers is the average of the diffusion coefficients of free and buffer-bound Ca2+, weighted by the buffer capacities:

Deff=DCa+DEκE+DNκN1+κE+κN

Here, DCa, DE, and DN are the diffusion coefficients of free Ca2+, endogenous, and native buffers in the cytoplasm. Thus, the presence of added buffer may break down normal diffusional barriers in small cellular compartments that are critical for physiological Ca2+ signaling.

In summary, in response to a brief rise in intracellular Ca2+ following synaptic activity or a back-propagating action potential, the presence of Ca2+ buffers reduces the magnitude of the increase in free Ca2+, prolongs the kinetics of decay of the Ca2+ transient, and increases the diffusional spread of Ca2+.

A second complication that arises in the study of synaptic Ca2+ signaling is that dendritic spines are small (Harris and Stevens 1988, 1989; Südhof and Rizo 2011) Thus, many features, including the spine neck and PSD, cannot be accurately imaged with visible light-based microscopy. In addition, mammalian brain tissue is highly scattering and absorbent to visible photons, rendering the direct visualization of dendritic spines difficult. For this reason, much of the recent study of Ca2+ signaling in dendritic spines has taken advantage of two-photon laser-scanning fluorescence microscopy. This approach uses the nonlinear near-infrared light-based excitation of fluorophores to allow imaging of dendritic spines deep within brain tissue (Denk et al. 1990; Denk and Svoboda 1997). However, it does so at the cost of reduced imaging resolution (owing to the longer wavelengths used), further obscuring many features of dendritic spines (however, for a practical consideration of this point, see Cox and Sheppard 2004). Importantly, the advent of laser-scanning supra-resolution microscopy, in which imaging resolution is better than the limit imposed by the diffraction of light, may soon overcome this challenge.

Last, the study of synaptic Ca2+ signaling in dendritic spines is made difficult by the complexity of the mammalian brain. The high packing density of synapses and the dense intercrossing of axons within the neuropil make it difficult to stimulate a single synapse in isolation. It is even more challenging to visualize a dendritic spine while selectively stimulating the apposed presynaptic axon (although, see Yuste and Denk 1995; Emptage et al. 1999; Oertner et al. 2002). Thus, the study of postsynaptic signaling in dendritic spines was greatly helped by the development of photosensitive derivatives of neurotransmitters that are inert in their parent form but release a fully functional neurotransmitter following exposure to light of the appropriate wavelength. The best characterized of these is MNI-glutamate, which can be readily photolyzed via two-photon excitation to selectively activate glutamate receptors on individual visualized spines within complex brain tissue (Canepari et al. 2001; Matsuzaki et al. 2001). This approach has been coupled with two-photon fluorescence imaging to allow the direct probing of synaptic Ca2+ signaling in an individual dendritic spine without the need to electrically activate axons.

MECHANISMS OF Ca2+ ENTRY

Synaptically coupled Ca2+ signals within dendritic spines originate from several sources that vary in the magnitude of Ca2+ transient produced, the time course of Ca2+ rise and decay, and the coupling to downstream biochemical pathways. Here, we consider the three most well-characterized sources of synaptic Ca2+: ionotropic glutamate receptors, voltage-gated calcium channels, and internal calcium stores.

Glutamate Receptors

Ionotropic glutamate receptors comprise a heterogeneous group of ligand-gated ion channels, which bind the amino acid glutamate and are permeable to monovalent (and, in some cases, divalent) cations. They show reversal potentials near 0 mV and mediate the vast majority of excitatory synaptic transmission in the nervous system. Glutamate receptors were initially grouped by their sensitivity to exogenous agonists, including N-methyl-d-aspartate (NMDA), 2-amino-3-(5-methyl-3-oxo-1,2-oxazol-4-yl) propanoic acid (AMPA), and kainic acid. This broad classification—into NMDA-, AMPA-, and kainate-type glutamate receptors—has remained useful and is now known to reflect different subunit compositions. As discussed below, glutamate receptors are found throughout the central nervous system on dendritic shafts and spines of multiple cell types and contribute to cytosolic Ca2+ elevation both directly and indirectly.

NMDA-Type Glutamate Receptors

NMDA-type glutamate receptors (NMDARs) show the highest fractional Ca2+ permeability of all glutamate receptors. Approximately 15% of the current through NMDARs is mediated by Ca2+ influx under physiological conditions (Schneggenburger et al. 1993; Jane et al. 2009). Thus, NMDARs are the predominant source of synaptic Ca2+ signals in a variety of cells, including pyramidal neurons in both the CA1 (Müller and Connor 1991; Regehr and Tank 1992; Mainen et al. 1999; Yuste et al. 1999; Kovalchuk et al. 2000; Sobczyk et al. 2005; Bloodgood and Sabatini 2007b) and CA3 (Reid et al. 2001) regions of the hippocampus, spiny stellate (Nevian and Sakmann 2004), and pyramidal neurons of the neocortex (Koester and Sakmann 1998; Schiller et al. 1998), striatal medium spiny neurons (Carter and Sabatini 2004; Higley and Sabatini 2010), and olfactory granule cells (Egger et al. 2005). Because NMDARs are usually located on dendritic spines, their activation by glutamate produces a highly compartmentalized Ca2+ transient that is largely limited to the activated spine, and blocking NMDAR activation with pharmacological antagonists such as APV or CPP typically reduces or eliminates synaptic Ca2+ signals in spines (Sabatini et al. 2002; Sobczyk et al. 2005). This rise in Ca2+ concentration regulates diverse processes including local biochemical signaling, protein/membrane trafficking, synaptic plasticity, and cell growth.

The number of NMDARs activated in a single spine during a synaptic event was estimated using two-photon fluorescent imaging (Nimchinsky et al. 2004). Results suggested that less than five and, in some cases, only a single receptor opens under physiological conditions. In a resting neuron, this opening produces a total Ca2+ influx of about 6000 ions into a spine head with a volume of 1 fL, corresponding to a concentration of ∼10 µm (Sabatini et al. 2002). Of course, 90–95% of these ions are rapidly bound by various internal buffering molecules, leading to a change in free Ca2+ concentration of ∼1 µm (Sabatini et al. 2002; Higley and Sabatini 2008).

Structurally, NMDARs are heteromeric tetramers typically consisting of two obligatory GluN1 subunits and two accessory GluN2 subunits (see Smart and Paoletti 2012). Some subunits also have multiple isoforms and, in some cases, multiple splice variants (Nakanishi et al. 1992; Sugihara et al. 1992). The structural diversity of NMDAR subunits contributes to functional heterogeneity of the channel. For example, in comparison with GluN2A-containing receptors, NMDARs with GluN2B subunits show higher affinity for glutamate binding, slower channel kinetics, and higher fractional Ca2+ permeability (Monyer et al. 1994; Vicini 1998; Sobczyk et al. 2005).

The influence of subunit composition on Ca2+ signaling suggests that activation of receptors composed of distinct subunit combinations may trigger different biological pat hways. This hypothesis is supported by immunogold electron microscopy studies showing that GluN2A- and GluN2B-containing receptors may be differentially expressed across synapses (He et al. 1998). Similarly, stimulation of single postsynaptic contacts with two-photon glutamate uncaging has shown that the contributions of GluN2A- and GluN2B-containing receptors to NMDAR-dependent currents and Ca2+ transients vary widely from spine to spine (Sobczyk et al. 2005). Antagonism of GluN2B-mediated Ca2+ transients with the selective blocker ifenprodil reduces both the amplitude and interspine variability of NMDAR-mediated Ca2+ transients, consistent with heterogeneous expression of GluN2B subunits at subsets of synapses (Sobczyk et al. 2005).

Distinct subunits may also play a critical role in the plasticity of synaptic strength. In many brain areas, there is a developmental switch in synapses from GluN2B- to GluN2A-containing NMDARs that coincides with the maturation of neuronal circuits (Monyer et al. 1994; Sheng et al. 1994). Additionally, subunit composition may be rapidly regulated in response to plasticity-inducing stimuli. Thus, induction of long-term potentiation at CA3 to CA1 synapses in the hippocampus of young rats is accompanied by a subunit switch from GluN2B to GluN2A (Bellone and Nicoll 2007). Finally, differential coupling to downstream Ca2+-dependent signaling pathways may allow GluN2A- versus GluN2B-containing receptors to have different functional implications for plasticity induction (Liu et al. 2004; Massey et al. 2004).

One of the most notable features of the NMDAR is that the conductance of cations, including Ca2+, is strongly regulated by membrane potential due to pore blockade by extracellular magnesium ions. Depolarization of the membrane potential by 20 mV decreases the affinity of Mg2+ for the NMDAR by nearly 10-fold (Jahr and Stevens 1990b). This property allows NMDARs to serve as coincidence detectors, in that synaptic current and local Ca2+ influx are strongly augmented by near-simultaneous membrane depolarization and glutamate binding. Postsynaptic membrane depolarization that relieves Mg2+ block occurs following strong synaptic activation. In addition, transient depolarization occurs when action potentials propagate antidromically through the dendritic arbor, potentiating Ca2+ influx through NMDARs (Yuste and Denk 1995; Magee and Johnston 1997; Koester and Sakmann 1998; Carter and Sabatini 2004; Nevian and Sakmann 2004). This enhancement may contribute to Ca2+-dependent synaptic plasticity that is mediated by precisely timed presynaptic and postsynaptic activity (Bender et al. 2006; Nevian and Sakmann 2006). Nevertheless, even at the resting potentials of most cells, Mg2+ block is incomplete, and glutamate binding to NMDARs can evoke Ca2+ influx in the absence of additional depolarization (Jahr and Stevens 1990a; Sabatini et al. 2002).

Ca2+ influx through NMDARs is also regulated by receptor phosphorylation, providing a biochemical means to alter synaptic Ca2+ signals. Protein kinase A (PKA) regulates the Ca2+ permeability of both GluN2A- and GluN2B-containing receptors (Skeberdis et al. 2006; Higley and Sabatini 2008; Chalifoux and Carter 2010). This mechanism underlies the modulation of NMDAR Ca2+ signaling by various G-protein-coupled receptors (see Box 2). Ca2+ permeability of NMDARs is also controlled by a negative-feedback loop such that repetitive activation of GluN2B-containing receptors activates a serine-threonine phosphatase that decreases Ca2+ permeability (Sobczyk and Svoboda 2007).

BOX 2. REGULATION OF NMDA RECEPTOR-DEPENDENT Ca2+ INFLUX BY PKA.

Ca2+ influx through NMDA-type glutamate receptors (NMDARs) is an essential step in the linkage between synaptic transmission and a variety of cellular processes including synaptogenesis, long-term changes in synaptic efficacy, membrane protein trafficking, gene transcription and translation, and cell death and survival pathways (Kennedy et al. 2005). Dysregulation of NMDAR-mediated Ca2+ influx is implicated in schizophrenia and in excitotoxic cell death associated with stroke, epilepsy, head trauma, and neurodegenerative disease (Lau and Zukin 2007; Lau and Tymianski 2010). Intriguingly, several recent studies have shown that the Ca2+ permeability of NMDARs is controlled by distinct neuromodulatory systems in the brain, opening up new avenues in our understanding of this signaling pathway.

NMDARs are regulated by the activity of protein kinase A (PKA), which, along with protein phosphatase-1, is coupled to NMDARs via an A-kinase anchoring protein (AKAP) (Westphal et al. 1999). Furthermore, NMDARs are known molecular targets of PKA phosphorylation. Skeberdis et al. (2006) showed that PKA selectively augmented the permeation of Ca2+ ions through NMDARs in both cultured hippocampal neurons and acutely prepared hippocampal slices. This modulation was further seen as an enhancement of NMDAR-mediated Ca2+ transients in synaptically activated dendritic spines with no concomitant change in total synaptic current. Blocking PKA activity reduced the early phase of NMDAR-dependent long-term potentiation of hippocampal synapses, suggesting that NMDAR Ca2+ permeability is a key target for the modulation of synaptic plasticity.

Higley and Sabatini (2010) showed that the activation of type 2 dopamine receptors (D2Rs) inhibited NMDAR-mediated synaptic Ca2+ influx into the dendritic spines of striatal medium spiny neurons. D2Rs are negatively coupled to cyclic AMP generation and PKA activity via their associated Gαi subunit, and the inhibition was both mimicked and occluded by blockers of PKA. Moreover, the actions of D2Rs were opposed by activating type A2 adenosine receptors (A2ARs), which are positively coupled to PKA via Gαs. These results suggest a likely cellular mechanism underlying the analogous actions of D2Rs and A2ARs on NMDAR-dependent long-term plasticity of corticostriatal synapses (Shen et al. 2008). In addition, Chalifoux and Carter (2010) found that activation of Gαi-coupled GABAB receptors also inhibits synaptic NMDAR-mediated Ca2+ influx in pyramidal neurons of the neocortex, further indicating the generalization of this mechanism.

These studies provide strong links between the activity of PKA and the functional regulation of synaptic Ca2+ signaling. Moreover, they provide a mechanism for coupling neuromodulatory pathways to myriad cellular phenomenon such as long-term synaptic plasticity. Given the connection between NMDAR-mediated Ca2+ signals and neuropsychiatric disease, they suggest new avenues for potential therapeutic interventions.

Non-NMDA-Type Glutamate Receptors

AMPA-type glutamate receptors (AMPARs) are heteromeric tetramers, typically comprising dimer pairs of GluA2 and either GluA1, GluA3, or GluA4 (Greger et al. 2007; Nakagawa 2010). The presence of a GluA2 subunit renders AMPARs minimally permeable to Ca2+ ions. However, a subset of AMPARs lacking GluA2 subunits is expressed in populations of inhibitory neurons, including cortical and hippocampal interneurons (Burnashev et al. 1992; Cull-Candy et al. 2006), striatal medium spiny neurons (Carter and Sabatini 2004), and cerebellar Purkinje cells (Denk et al. 1995). Ca2+-permeable AMPARs have also been found in excitatory neurons of the hippocampus (Thiagarajan et al. 2005; Plant et al. 2006) and amygdala (Clem and Huganir 2010). The presence of GluA2-lacking AMPARs is often identified by their characteristic electrophysiological signature of strong inward rectification at depolarized membrane potentials. Because of their fast kinetics, Ca2+ influx through these receptors is significantly briefer and of lower magnitude than that occurring through NMDARs. Furthermore, the functional consequences of Ca2+ influx through GluA2-lacking AMPARs is less clear, although they may contribute to synaptic plasticity in some cells (Thiagarajan et al. 2005; Plant et al. 2006; Clem and Huganir 2010).

Importantly, AMPARs indirectly contribute to Ca2+ signaling by providing membrane depolarization, relieving Mg block from NMDARs, and activating voltage-gated Ca2+ channels (see below). Similarly, kainate-type glutamate receptors are minimally permeable to Ca2+. However, they may also contribute indirectly to Ca2+ influx via membrane depolarization in some cells (Jane et al. 2009).

Voltage-Gated Ca2+ Channels

Voltage-gated Ca2+ channels (VGCCs) make up a second class of contributors to Ca2+ influx. VGCCs are heteromeric complexes comprising a primary α1 pore-forming subunit and four additional associated subunits (α2, β, or γ). The molecular identity of the α1 subunit determines the functional classification of neuronal channels as L-type (α1C and D or CaV1.2 and 1.3), P/Q-type (α1A or CaV2.1), N-type (α1B or CaV2.2), R-type (α1E or CaV2.3), or T-type (α1G, H, and I or CaV3.1, 3.2, and 3.3) (Hille 2001).

Within dendrites and spines, VGCCs open following synaptically evoked depolarization (Miyakawa et al. 1992; Christie et al. 1995; Denk et al. 1995; Eilers et al. 1995; Magee et al. 1995; Markram et al. 1995; Finch and Augustine 1998; Schiller et al. 1998; Reid et al. 2001). Sufficient depolarization for VGCC activation can also be provided by the back-propagation of somatically generated action potentials that spread antidromically through at least the proximal portions of the dendritic arbor (Callaway and Ross 1995; Schiller et al. 1995; Yuste and Denk 1995; Svoboda et al. 1997; Helmchen et al. 1999; Koester and Sakmann 2000; Waters et al. 2003; Carter and Sabatini 2004; Nevian and Sakmann 2004; Bloodgood and Sabatini 2007b). Opening of VGCCs by action potential-evoked depolarization was used, in combination with fluctuation analysis, to estimate the number of VGCCs in a single dendritic spine at about one to 20, with the number correlating with spine volume (Sabatini and Svoboda 2000).

Numerous electrophysiological and imaging studies have revealed considerable heterogeneity in the expression of VGCC classes between different dendritic regions and across different cell types and species. In CA1 neurons of the hippocampus, Ca2+ influx into the dendritic shaft occurs via L-, R-, and T-type channels (Christie et al. 1995; Magee et al. 1995; Sabatini and Svoboda 2000), whereas influx into individual spine heads appears to be primarily limited to R- and T-type channels with a small contribution from L-type channels (Sabatini and Svoboda 2000; Yasuda et al. 2003; Hoogland and Saggau 2004; Bloodgood and Sabatini 2007b). In cortical pyramidal neurons, dendritic VGCCs include L-, N-, P/Q-, and R-type channels (Markram et al. 1995), whereas L-, P/Q-, and T-type channels are found in spines (Koester and Sakmann 2000). T-type VGCCs also contribute to dendritic Ca2+ signals in both olfactory granule cells (Egger et al. 2005) and cerebellar Purkinje cells (Isope and Murphy 2005). In the lateral nucleus of the amygdala, cortical and thalamic synapses are associated with spines of different sizes and voltage-gated Ca2+ channel content such that thalamic inputs are preferentially found on large spines that contain R-type channels, which may underlie differences in the amplitude of evoked Ca2+ transients and the ability to express spike-timing dependent plasticity at these two classes of synapses (Humeau et al. 2005). Within striatal medium spiny neurons, Ca2+ influx occurs primarily through R- and T-type channels in both dendritic shafts and spines (Higley and Sabatini 2010). Nevertheless, interpretation of many of these studies is limited by imprecise mapping between pharmacological sensitivity and VGCC α-subunit expression, making the molecular composition of the channels mediating dendritic and spine Ca2+ influx difficult to establish conclusively. This is particular true for L-type channels (see Box 3).

BOX 3. MYSTERIES OF R- AND L-TYPE VOLTAGE-GATED Ca2+ CHANNELS—Ca2+ MICRODOMAIN ORGANIZATION OF THE SPINE.

Among the many sources of Ca2+ that are active in dendritic spines, the R- and L-type voltage-gated Ca2+ channels appear to have special functions that reveal an intricate subdivision of the spine into smaller Ca2+ signaling microdomains. L-type channels are often defined pharmacologically by their sensitivity to blockade by dihydropyridines such as nifedipine and nimodipine. However, neuronal L-types comprise two distinct molecular subclasses, CaV1.2 and CaV1.3, based on the identity of the pore-forming subunit, encoded by the α1C or α1D gene, respectively. One major challenge to analyzing the function of L-type channels is that CaV1.3/α1D -encoded channels have diminished sensitivity to dihydropyridines and require substantial depolarization to display significant block by these agents (Lipscombe et al. 2004).

These properties present difficulties in detecting the activation of L-type channels during brief physiological stimuli. For example, action potential-evoked Ca2+ influx in dendritic spines of striatal medium spiny neurons has little or no sensitivity to concentrations of dihydropyridines that are selective for blockade of L-type channels. However, prolonged step depolarizations show that dihydropyridine-sensitive Ca2+ channels are, in fact, present in the spine (Higley and Sabatini 2010). A similarly perplexing result was found by the study of Yasuda et al. (2003), which described that Ca2+ influx through L-type Ca2+ channels activated a kinase cascade producing inhibition of other spine Ca2+ channels. However, no L-type-mediated Ca2+ influx could be detected in either dendrites or spines via fluorescence imaging. The resolution of this paradox is likely that the total Ca2+ influx carried by L-type channels is too small to detect but that this Ca2+ has a privileged capacity to activate downstream signaling cascades. A close apposition of Ca-sensitive proteins to the mouth of the L-type channel would allow brief and large Ca2+ accumulation within a “microdomain.” Tethering of CaMKII to the carboxy-terminal domain of the α-subunit is one such example (Hudmon et al. 2005).

Another example of Ca2+ microdomain signaling involves R-type voltage-gated channels (α1E/CaV2.3). These are probably the least well-characterized class of voltage-gated Ca2+ channel because of a historical lack of sensitivity to known pharmacological agents (R = resistant). However, a spider toxin (SNX-482) is available that blocks these channels with good specificity, allowing the study of their contribution to Ca2+ signaling (Newcomb et al. 1998). Early reports showed that R-type channels likely contributed the bulk of action potential-evoked Ca2+ influx in spines of hippocampal pyramidal neurons (Sabatini and Svoboda 2000; Yasuda et al. 2003). Interestingly, recent studies using SNX-482 showed that blocking Ca2+ influx through these channels increased the amplitude of synaptic potentials and, quite perplexingly, also increased the magnitude of synaptically evoked Ca2+ transients in active spines (Bloodgood and Sabatini 2007b). To explain this observation, a series of studies has now shown that Ca2+ entering through R-type channels has a privileged ability to activate small conductance type Ca-activated potassium channels known as SK channels. SK channels open quickly during synaptic potentials and repolarize the active spine, thereby reducing the magnitude and of synaptic potentials and associated NMDA receptor-dependent Ca2+ influx (Ngo-Anh et al. 2005; Bloodgood and Sabatini 2007b). Blocking Ca2+ influx through R-type channels prevents the activation of SK channels, essentially disinhibiting synaptic signals. Interestingly, this signaling cascade is modulated by muscarinic cholinergic receptors, and the inhibition of SK channels explains much of the ability of muscarinic agents to enhance synaptic potentials, contributing to the induction of synaptic plasticity and hippocampal-dependent learning (Giessel and Sabatini 2010).

In summary, both L-type and R-type voltage-gated Ca2+ channels appear to have privileged functions in dendritic spines. That is, Ca2+ entering through these channels activates signaling cascades that are not otherwise triggered by bulk elevation in cytoplasmic Ca2+ concentration. These results are undoubtedly a manifestation of a Ca2+ microdomain organization within the spine in which specific Ca2+ sources are physically associated with Ca2+-sensitive proteins, producing important consequences for synaptic function.

Internal Stores

The contributions of Ca2+ release from internal stores to dendritic and spine Ca2+ transients following synaptic activation are more controversial and likely depend critically on the cell type under consideration as well as the experimental protocols used. Strong activation of glutamatergic inputs to hippocampal CA1 pyramidal neurons can lead to activation of group I metabotropic glutamate receptors (mGluRs), triggering a phospholipase C (PLC)– and inositol triphosphate (IP3)–dependent Ca2+ release from internal stores, and contributing to long-term heterosynaptic plasticity (Watanabe et al. 2006; Dudman et al. 2007; Hong and Ross 2007). This may lead to Ca2+ waves throughout large regions of the apical dendrite, regulating cell-wide signaling pathways. Putative IP3- and Ca2+-induced Ca2+ release from internal stores may also contribute to synaptically evoked Ca2+ transients in individual dendritic spines (Emptage et al. 1999). However, other studies have failed to find evidence for Ca2+ release from internal stores following more limited synaptic stimulation of hippocampal afferents (Mainen et al. 1999; Yuste et al. 1999; Kovalchuk et al. 2000). This disparity may be due to Ca2+ depletion of internal stores during whole-cell recordings (Hong and Ross 2007) as well as by the relatively few spines that contain endoplasmic reticulum (Holbro et al. 2009). More evidence for synaptically evoked Ca2+ release from internal stores exists for cerebellar Purkinje cells, where several groups have shown that activation of parallel fiber inputs can lead to Ca2+ release via a Group 1 mGluR-PLC-IP3-coupled pathway (Finch and Augustine 1998; Takechi et al. 1998; Miyata et al. 2000; Wang et al. 2000).

MECHANISMS OF Ca2+ HANDLING AND CLEARANCE

Following entry into the spine, free Ca2+ concentration increases and then decreases rapidly because of the action of Ca2+-binding proteins and Ca2+ extrusion mechanisms. This rapid clearance is essential to maintaining spatially and temporally localized Ca2+ signals that can mediate synapse-specific (homosynaptic) forms of plasticity as well as the induction of spike timing-dependent plasticity. The timing and magnitude of synaptic Ca2+ transients in active dendritic spines are determined by an interplay between the kinetics of opening of Ca2+ sources, the on and off rates of Ca2+ binding to proteins and lipids, and the membrane densities and transport rates of Ca2+ transporters and exchangers.

Ca2+-Binding Proteins

The human genome contains more than 200 proteins with EF-hands, prototypical Ca2+-binding domains, suggesting a large family of Ca2+-binding proteins. Some of these represent direct Ca2+-activated enzymes, such as the phosphatase calcineurin or the protease calpain. Others, such as calmodulin, are present at high levels and, upon Ca2+ binding, undergo a conformational change that promotes interaction with and activation of enzymes such as CaMKII. Notably, CaMKII also binds Ca2+ directly, leading to a synergistic activation via Ca2+ and Ca2+/calmodulin signaling. However, a subset of Ca2+-binding proteins may simply bind Ca2+ in order to modulate the spatial and temporal profile of Ca2+ signaling. This group includes the prototypical Ca2+ buffering proteins parvalbumin, calretinin, and calbindin. These proteins are found in the dendrites and axons of many neuron classes, particularly inhibitory interneurons. However, the properties and expression patterns of Ca2+ buffering proteins vary greatly, and their functional significance is poorly understood.

The Ca2+ buffering capacity of a molecule is typically referred to as κ and is roughly defined as the ratio of the number of Ca2+ ions that are bound to proteins to the number of Ca2+ ions that remain free (see Box 1) (Neher and Augustine 1992):

κ=[Ca]bound[Ca]free

The value of κ has profound implications for Ca2+ signaling because the amplitude of evoked Ca2+ transients is inversely proportional to κ, whereas their duration is directly proportional to κ (Box 1). For pyramidal neurons, κ is 100–200 near the soma and base of the apical dendrite, whereas κ = 20 in distal dendrites and dendritic spines. Thus, only ∼1%–5% of the Ca2+ that enters the cell remains unbound (Helmchen et al. 1996; Lee et al. 2000a; Maravall et al. 2000; Sabatini et al. 2002). The relatively low κ of spines and dendrites allows for rapid Ca2+ signaling in which large transients are generated when a Ca2+ channel opens and are rapidly dissipated (with ∼15 msec) after closure of the channel. For these reasons, action potential and synaptically evoked Ca2+ transients are large (∼1 µm) and closely follow the kinetics of opening of Ca2+ sources (Sabatini et al. 2002).

In contrast, κ is significantly larger in cerebellar Purkinje cells and inhibitory interneurons that express parvalbumin (PV), a kinetically slow and high-affinity Ca2+-binding protein (Lee et al. 2000a,b; Goldberg et al. 2003; Soler-Llavina and Sabatini 2006). The peculiar properties of PV shape Ca2+ signaling in these cells. Because PV binds Mg2+ and Ca2+ competitively, Ca2+ equilibrates slowly with PV such that the binding reaction does not reach equilibrium for tens of milliseconds. During this “non-equilibrium” phase of Ca2+ signaling, Ca2+ remains free and at elevated levels. However, the binding of Ca2+ to PV subsequently returns Ca2+ to near resting levels, and once Ca2+ reaches equilibrium with PV, the further decay of [Ca2+] to resting levels occurs slowly (Collin et al. 2005; Soler-Llavina and Sabatini 2006; Muller et al. 2007). For these reasons, in PV-expressing neurons, the decay of evoked Ca2+ transients is typically complex and displays multiple time constants. In contrast, in pyramidal neurons for which Ca2+ buffers are thought to rapidly bind Ca2+, equilibrium is reached quickly and the time course of Ca2+ clearance is well described by a single exponential process corresponding to extrusion (Helmchen et al. 1996; Sabatini et al. 2002).

In addition to regulating the time course of Ca2+ clearance, PV may play an important role in compartmentalizing Ca2+ in neurons that lack dendritic spines (Goldberg et al. 2003; Soler-Llavina and Sabatini 2006). This has been closely examined in cerebellar stellate cells, aspiny cells that express synapse-specific and Ca2+-dependent forms of plasticity. If Ca2+ diffused rapidly away from the site of entry into the dendrite, it would be impossible to maintain the synapse specificity of Ca2+-dependent processes. In contrast, if an aspiny neuron restricted diffusion by the expression of a high concentration of fast-acting Ca2+ buffer, then the amplitude of evoked Ca2+ transients might be too small to effectively activate Ca2+-dependent processing. PV allows for large Ca2+ transients because of its slow kinetics, but because of its high affinity, effectively buffers Ca2+ after a short lag. This allows synaptic Ca2+ in smooth dendrites to induce synapse specific Ca2+-dependent forms of plasticity (Soler-Llavina and Sabatini 2006).

The function of Ca2+-binding proteins in pyramidal neurons is less clear. Because the spine head is a small volume that is separated from the dendrite by a thin neck, it naturally allows for large and compartmentalized Ca2+ transients such that Ca2+-binding proteins are not needed for this purpose. Thus, Ca2+-binding proteins may largely serve a signaling role. This is illustrated by calmodulin, which binds four Ca2+ ions and once fully occupied can bind to and activate a variety of proteins, being a major determinant of the Ca2+ buffering capacity of dendritic spines (Kakiuchi et al. 1982; Sabatini et al. 2002), at least for CA1 pyramidal neurons. On the other hand, it is difficult to draw conclusions that apply to all pyramidal neurons because there are differences in the Ca2+ buffering across these neurons that are of functional importance. For example, pyramidal neurons of the CA2 subfield of the hippocampus express more calcium-binding proteins than CA1 pyramidal neurons (Leranth and Ribak 1991; Seress et al. 1993; Lein et al. 2007), a difference the reduces the amplitude of spine Ca2+ transients (see below) (Simons et al. 2009). The relative low amplitude of synaptic Ca2+ transients in active spines of these cells may explain the difficulty of induction of long-term potentiation in CA2 compared with in CA1 as well as their resistance to seizure-induced death (Leranth and Ribak 1991; Simons et al. 2009). It is interesting to note that the high κ is reflected in the low amplitude of evoked transients but that a high Ca2+ extrusion rate (see below) compensates for the expected slowing of Ca2+ clearance, yielding Ca2+ transients whose kinetics do not differ significantly from those in CA1 (Simons et al. 2009).

Ca2+ Extrusion

Three main avenues exist to clear Ca2+ from the cytoplasm. First, ATP-dependent Ca2+ pumps, such as the plasmalemmal Ca2+ ATPase (PMCA), move Ca2+ across the cell membrane from the cytosol into the extracellular space. PMCA isoforms have been found both in the postsynaptic density and localized to spines, placing them appropriately to rapidly clear Ca2+ from the spine head (Garside et al. 2009; Burette et al. 2010; Kenyon et al. 2010). The high expression of PMCA in pyramidal neurons may explain the rapid Ca2+ clearance capacity of these cells (Jensen et al. 2004; Kip et al. 2006). The efficacy of PMCA is itself regulated by intracellular Ca2+ such that at high Ca2+ concentrations, the efficiency of PMCA-dependent Ca2+ extrusion is decreased, resulting in slowed Ca2+ clearance during prolonged excitation (Scheuss et al. 2006). The mechanism of this slowing may include Ca2+-dependent activation of proteases that act on PMCA (Ferragamo et al. 2009), although this irreversible mechanism cannot account for the transient slowing of Ca2+ clearance seen following action potential trains (Scheuss et al. 2006). The proteins comprising the Na+/Ca2+ exchangers (NCX1-3) have also been localized to dendrites and dendritic spines with the distribution varying by isoform (Lorincz et al. 2007; Minelli et al. 2007). Functional analysis suggests that these proteins play a major functional role in clearance of spine Ca2+ (Scheuss et al. 2006).

Second, Ca2+ can be actively pumped into the lumen of intracellular organelles. The sarco-endoplasmic reticulum Ca-ATPase (SERCA) moves Ca2+ across the endoplasmic reticulum (ER) membrane and sequesters it within this organelle. The activity of SERCA may account for up to 50% of the clearance of spine Ca2+ transients during single action or synaptic potentials (Majewska et al. 2000a; Sabatini et al. 2002). This same pump serves to constitutively load ER with Ca2+, and its activity is necessary to support IP3- and Ca2+-induced Ca2+ release from intracellular stores (see “Internal Stores” above). Although mitochondria can also rapidly uptake and release Ca2+, the significance of mitochondrial Ca2+ handling to postsynaptic signaling is unclear. Pharmacological manipulation of mitochondrial Ca2+ signaling is typically accomplished by destroying the mitochondrial membrane potential and associated ATP production. Thus, it has been difficult to selectively examine the function of mitochondrial Ca2+ uptake without also perturbing the energetic state of the cell.

Third, in theory, Ca2+ can diffuse away from the site of entry. This mechanism of clearance can only be relevant in conditions in which gradients of Ca2+ exist within the cell, such as when a synapse is active and its neighbors are not. Within the spine head, Ca2+ is expected to quickly reach diffusional equilibration such that the concentration of free Ca2+ throughout the head is likely uniform within ∼1 msec of closure of the Ca2+ sources. Synaptically evoked Ca2+ signals do form a gradient between the spine head and dendritic shaft, suggesting that Ca2+ might dissipate by diffusion across the spine neck (Majewska et al. 2000a,b; Noguchi et al. 2005). However, other studies suggest that when the effects of Ca2+ indicators on Ca2+ signaling are properly accounted for, free Ca2+ has a very short lifetime in the spine and does not diffuse far (Sabatini et al. 2002; Sobczyk et al. 2005). Instead, Ca2+ is quickly removed from the spine cytoplasm before it has a chance to diffuse across the neck.

CONCLUSION

In summary, synaptically evoked Ca2+ transients in dendritic spines are shaped by a complex interplay between many Ca2+ sources, pumps, and binding proteins, as well as by the biophysics of Ca2+ diffusion and spine morphology. Many classes of intracellular proteins respond to changes in postsynaptic Ca2+ to transduce electrical signaling into the regulation of enzymatic cascades. Our understanding of the factors governing Ca2+ signaling has increased dramatically, but many basic questions, such as the importance of intracellular Ca2+ stores, remain unanswered. Furthermore, as more classes of neurons are studied biophysically, it is becoming clear that neurons have highly heterogeneous Ca2+ handling mechanisms that likely have significant functional implications.

Footnotes

Editors: Morgan Sheng, Bernardo Sabatini, and Thomas C. Südhof

Additional Perspectives on The Synapse available at www.cshperspectives.org

REFERENCES

*Reference is also in this collection.

  1. Bellone C, Nicoll RA 2007. Rapid bidirectional switching of synaptic NMDA receptors. Neuron 55: 779–785 [DOI] [PubMed] [Google Scholar]
  2. Bender VA, Bender KJ, Brasier DJ, Feldman DE 2006. Two coincidence detectors for spike timing-dependent plasticity in somatosensory cortex. J Neurosci 19: 4166–4177 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Bloodgood BL, Sabatini BL 2007a. Ca2+ signaling in dendritic spines. Curr Opin Neurobiol 17: 345–351 [DOI] [PubMed] [Google Scholar]
  4. Bloodgood BL, Sabatini BL 2007b. Nonlinear regulation of unitary synaptic signals by CaV(2.3) voltage-sensitive calcium channels located in dendritic spines. Neuron 53: 249–260 [DOI] [PubMed] [Google Scholar]
  5. Bloodgood BL, Giessel AJ, Sabatini BL 2009. Biphasic synaptic Ca influx arising from compartmentalized electrical signals in dendritic spines. PLoS Biol 7: e1000190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Burette AC, Strehler EE, Weinberg RJ 2010. A plasma membrane Ca2+ ATPase isoform at the postsynaptic density. Neuroscience 169: 987–993 [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Burnashev N, Monyer H, Seeburg PH, Sakmann B 1992. Divalent ion permeability of AMPA receptor channels is dominated by the edited form of a single subunit. Neuron 8: 189–198 [DOI] [PubMed] [Google Scholar]
  8. Callaway JC, Ross WN 1995. Frequency-dependent propagation of sodium action potentials in dendrites of hippocampal CA1 pyramidal neurons. J Neurophysiol 74: 1395–1403 [DOI] [PubMed] [Google Scholar]
  9. Canepari M, Nelson L, Papageorgiou G, Corrie JE, Ogden D 2001. Photochemical and pharmacological evaluation of 7-nitroindolinyl-and 4-methoxy-7-nitroindolinyl-amino acids as novel, fast caged neurotransmitters. J Neurosci Methods 112: 29–42 [DOI] [PubMed] [Google Scholar]
  10. Carter AG, Sabatini BL 2004. State-dependent calcium signaling in dendritic spines of striatal medium spiny neurons. Neuron 74: 483–493 [DOI] [PubMed] [Google Scholar]
  11. Chalifoux JR, Carter AG 2010. GABAB receptors modulate NMDA receptor calcium signals in dendritic spines. Neuron 66: 101–113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Christie BR, Eliot LS, Ito K, Miyakawa H, Johnston D 1995. Different Ca2+ channels in soma and dendrites of hippocampal pyramidal neurons mediate spike-induced Ca2+ influx. J Neurophysiol 73: 2553–2557 [DOI] [PubMed] [Google Scholar]
  13. Clem RL, Huganir RL 2010. Calcium-permeable AMPA receptor dynamics mediate fear memory erasure. Science 330: 1108–1112 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Collin T, Chat M, Lucas MG, Moreno H, Racay P, Schwaller B, Marty A, Llano I 2005. Developmental changes in parvalbumin regulate presynaptic Ca2+ signaling. J Neurosci 25: 96–107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Cox G, Sheppard CJR 2004. Practical limits of resolution in confocal and non-linear microscopy. Microsc Res Tech 63: 18–22 [DOI] [PubMed] [Google Scholar]
  16. Cull-Candy S, Kelly L, Farrant M 2006. Regulation of Ca2+-permeable AMPA receptors: Synaptic plasticity and beyond. Curr Opin Neurobiol 16: 288–297 [DOI] [PubMed] [Google Scholar]
  17. Denk W, Svoboda K 1997. Photon upmanship: Why multiphoton imaging is more than a gimmick. Neuron 18: 351–357 [DOI] [PubMed] [Google Scholar]
  18. Denk W, Strickler JH, Webb WW 1990. Two-photon laser scanning fluorescence microscopy. Science 248: 73–76 [DOI] [PubMed] [Google Scholar]
  19. Denk W, Sugimori M, Llinás R 1995. Two types of calcium response limited to single spines in cerebellar Purkinje cells. Proc Natl Acad Sci 92: 8279–8282 [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Dudman JT, Tsay D, Siegelbaum SA 2007. A role for synaptic inputs at distal dendrites: Instructive signals for hippocampal long-term plasticity. Neuron 56: 866–879 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Egger V, Svoboda K, Mainen ZF 2005. Dendrodendritic synaptic signals in olfactory bulb granule cells: Local spine boost and global low-threshold spike. J Neurosci 25: 3521–3530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Eilers J, Augustine GJ, Konnerth A 1995. Subthreshold synaptic Ca2+ signalling in fine dendrites and spines of cerebellar Purkinje neurons. Nature 373: 155–158 [DOI] [PubMed] [Google Scholar]
  23. Emptage N, Bliss TV, Fine A 1999. Single synaptic events evoke NMDA receptor-mediated release of calcium from internal stores in hippocampal dendritic spines. Neuron 22: 115–124 [DOI] [PubMed] [Google Scholar]
  24. Ferragamo MJ, Reinardy JL, Thayer SA 2009. Ca2+-dependent, stimulus-specific modulation of the plasma membrane Ca2+ pump in hippocampal neurons. J Neurophysiol 101: 2563–2571 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Finch EA, Augustine GJ 1998. Local calcium signalling by inositol-1,4,5-trisphosphate in Purkinje cell dendrites. Nature 396: 753–756 [DOI] [PubMed] [Google Scholar]
  26. Garside ML, Turner PR, Austen B, Strehler EE, Beesley PW, Empson RM 2009. Molecular interactions of the plasma membrane calcium ATPase 2 at pre- and post-synaptic sites in rat cerebellum. Neuroscience 162: 383–395 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Giessel AJ, Sabatini BL 2010. M1 muscarinic receptors boost synaptic potentials and calcium influx in dendritic spines by inhibiting postsynaptic SK channels. Neuron 68: 936–947 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Goldberg JH, Tamas G, Aronov D, Yuste R 2003. Calcium microdomains in aspiny dendrites. Neuron 40: 807–821 [DOI] [PubMed] [Google Scholar]
  29. Greger IH, Ziff EB, Penn AC 2007. Molecular determinants of AMPA receptor subunit assembly. Trends Neurosci 30: 407–416 [DOI] [PubMed] [Google Scholar]
  30. Grunditz A, Holbro N, Tian L, Zuo Y, Oertner TG 2008. Spine neck plasticity controls postsynaptic calcium signals through electrical compartmentalization. J Neurosci 28: 13457–13466 [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Harris KM, Stevens JK 1988. Dendritic spines of rat cerebellar Purkinje cells: Serial electron microscopy with reference to their biophysical characteristics. J Neurosci 8: 4455–4469 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Harris KM, Stevens JK 1989. Dendritic spines of CA 1 pyramidal cells in the rat hippocampus: Serial electron microscopy with reference to their biophysical characteristics. J Neurosci 9: 2982–2997 [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. *.Harris KM, Weinberg RJ 2012. Ultrastructure of synapses in the mammalian brain. Cold Spring Harb Perspect Biol 10.1101/cshperspect.a005587 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. He Y, Janssen WG, Morrison JH 1998. Synaptic coexistence of AMPA and NMDA receptors in the rat hippocampus: A postembedding immunogold study. J Neurosci Res 54: 444–449 [DOI] [PubMed] [Google Scholar]
  35. Helmchen F, Imoto K, Sakmann B 1996. Ca2+ buffering and action potential-evoked Ca2+ signaling in dendrites of pyramidal neurons. Biophys J 70: 1069–1081 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Helmchen F, Svoboda K, Denk W, Tank DW 1999. In vivo dendritic calcium dynamics in deep-layer cortical pyramidal neurons. Nat Neurosci 2: 989–996 [DOI] [PubMed] [Google Scholar]
  37. Higley MJ, Sabatini BL 2008. Calcium signaling in dendrites and spines: Practical and functional considerations. Neuron 59: 902–913 [DOI] [PubMed] [Google Scholar]
  38. Higley MJ, Sabatini BL 2010. Competitive regulation of synaptic Ca2+ influx by D2 dopamine and A2A adenosine receptors. Nat Neurosci 13: 958–966 [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Hille B 2001. Ion channels of excitable membranes, 3rd ed Sinauer Associates, Sunderland, MA [Google Scholar]
  40. Holbro N, Grunditz A, Oertner TG 2009. Differential distribution of endoplasmic reticulum controls metabotropic signaling and plasticity at hippocampal synapses. Proc Natl Acad Sci 106: 15055–15060 [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Hong M, Ross WN 2007. Priming of intracellular calcium stores in rat CA1 pyramidal neurons. J Physiol (Lond) 584: 75–87 [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Hoogland TM, Saggau P 2004. Facilitation of L-type Ca2+ channels in dendritic spines by activation of β2 adrenergic receptors. J Neurosci 24: 8416–8427 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Hudmon A, Schulman H, Kim J, Maltez JM, Tsien RW, Pitt GS 2005. CaMKII tethers to L-type Ca2+ channels, establishing a local and dedicated integrator of Ca2+ signals for facilitation. J Cell Biol 171: 537–547 [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Humeau Y, Herry C, Kemp N, Shaban H, Fourcaudot E, Bissière S, Lüthi A 2005. Dendritic spine heterogeneity determines afferent-specific Hebbian plasticity in the amygdala. Neuron 45: 119–131 [DOI] [PubMed] [Google Scholar]
  45. Isope P, Murphy TH 2005. Low threshold calcium currents in rat cerebellar Purkinje cell dendritic spines are mediated by T-type calcium channels. J Physiol (Lond) 562: 257–269 [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Jahr CE, Stevens CF 1990a. A quantitative description of NMDA receptor-channel kinetic behavior. J Neurosci 10: 1830–1837 [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Jahr CE, Stevens CF 1990b. Voltage dependence of NMDA-activated macroscopic conductances predicted by single-channel kinetics. J Neurosci 10: 3178–3182 [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Jane DE, Lodge D, Collingridge GL 2009. Kainate receptors: Pharmacology, function and therapeutic potential. Neuropharmacology 56: 90–113 [DOI] [PubMed] [Google Scholar]
  49. Jensen TP, Buckby LE, Empson RM 2004. Expression of plasma membrane Ca2+ ATPase family members and associated synaptic proteins in acute and cultured organotypic hippocampal slices from rat. Brain Res Dev Brain Res 152: 129–136 [DOI] [PubMed] [Google Scholar]
  50. Kakiuchi S, Yasuda S, Yamazaki R, Teshima Y, Kanda K, Kakiuchi R, Sobue K 1982. Quantitative determinations of calmodulin in the supernatant and particulate fractions of mammalian tissues. J Biochem 92: 1041–1048 [DOI] [PubMed] [Google Scholar]
  51. Kao JP 1994. Practical aspects of measuring [Ca2+] with fluorescent indicators. Methods Cell Biol 40: 155–181 [DOI] [PubMed] [Google Scholar]
  52. Kennedy MB, Beale HC, Carlisle HJ, Washburn LR 2005. Integration of biochemical signalling in spines. Nat Rev Neurosci 6: 423–434 [DOI] [PubMed] [Google Scholar]
  53. Kenyon KA, Bushong EA, Mauer AS, Strehler EE, Weinberg RJ, Burette AC 2010. Cellular and subcellular localization of the neuron-specific plasma membrane calcium ATPase PMCA1a in the rat brain. J Comp Neurol 518: 3169–3183 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Kip SN, Gray NW, Burette A, Canbay A, Weinberg RJ, Strehler EE 2006. Changes in the expression of plasma membrane calcium extrusion systems during the maturation of hippocampal neurons. Hippocampus 16: 20–34 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Koester HJ, Sakmann B 1998. Calcium dynamics in single spines during coincident pre- and postsynaptic activity depend on relative timing of back-propagating action potentials and subthreshold excitatory postsynaptic potentials. Proc Natl Acad Sci 95: 9596–9601 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Koester HJ, Sakmann B 2000. Calcium dynamics associated with action potentials in single nerve terminals of pyramidal cells in layer 2/3 of the young rat neocortex. J Physiol 529: 625–646 [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Kovalchuk Y, Eilers J, Lisman J, Konnerth A 2000. NMDA receptor-mediated subthreshold Ca2+ signals in spines of hippocampal neurons. J Neurosci 20: 1791–1799 [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Lau A, Tymianski M 2010. Glutamate receptors, neurotoxicity and neurodegeneration. Pflugers Arch 460: 525–542 [DOI] [PubMed] [Google Scholar]
  59. Lau CG, Zukin RS 2007. NMDA receptor trafficking in synaptic plasticity and neuropsychiatric disorders. Nat Rev Neurosci 8: 413–426 [DOI] [PubMed] [Google Scholar]
  60. Lee SH, Rosenmund C, Schwaller B, Neher E 2000a. Differences in Ca2+ buffering properties between excitatory and inhibitory hippocampal neurons from the rat. J Physiol 525: 405–418 [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Lee SH, Schwaller B, Neher E 2000b. Kinetics of Ca2+ binding to parvalbumin in bovine chromaffin cells: Implications for [Ca2+] transients of neuronal dendrites. J Physiol 525: 419–432 [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Lein ES, Hawrylycz MJ, Ao N, Ayres M, Bensinger A, Bernard A, Boe AF, Boguski MS, Brockway KS, Byrnes EJ, et al. 2007. Genome-wide atlas of gene expression in the adult mouse brain. Nature 445: 168–176 [DOI] [PubMed] [Google Scholar]
  63. Leranth C, Ribak CE 1991. Calcium-binding proteins are concentrated in the CA2 field of the monkey hippocampus: A possible key to this region’s resistance to epileptic damage. Exp Brain Res 85: 129–136 [DOI] [PubMed] [Google Scholar]
  64. Lipscombe D, Helton TD, Xu W 2004. L-type calcium channels: The low down. J Neurophysiol 92: 2633–2641 [DOI] [PubMed] [Google Scholar]
  65. Liu L, Wong TP, Pozza MF, Lingenhoehl K, Wang Y, Sheng M, Auberson YP, Wang YT 2004. Role of NMDA Receptor subtypes in governing the direction of hippocampal synaptic plasticity. Science 304: 1021–1024 [DOI] [PubMed] [Google Scholar]
  66. Lorincz A, Rozsa B, Katona G, Vizi ES, Tamas G 2007. Differential distribution of NCX1 contributes to spine-dendrite compartmentalization in CA1 pyramidal cells. Proc Natl Acad Sci 104: 1033–1038 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Magee JC, Johnston D 1997. A synaptically controlled, associative signal for Hebbian plasticity in hippocampal neurons. Science 275: 209–213 [DOI] [PubMed] [Google Scholar]
  68. Magee JC, Christofi G, Miyakawa H, Christie B, Lasser-Ross N, Johnston D 1995. Subthreshold synaptic activation of voltage-gated Ca2+ channels mediates a localized Ca2+ influx into the dendrites of hippocampal pyramidal neurons. J Neurophysiol 74: 1335–1342 [DOI] [PubMed] [Google Scholar]
  69. Mainen ZF, Malinow R, Svoboda K 1999. Synaptic calcium transients in single spines indicate that NMDA receptors are not saturated. Nature 399: 151–155 [DOI] [PubMed] [Google Scholar]
  70. Majewska A, Brown E, Ross J, Yuste R 2000a. Mechanisms of calcium decay kinetics in hippocampal spines: Role of spine calcium pumps and calcium diffusion through the spine neck in biochemical compartmentalization. J Neurosci 20: 1722–1734 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Majewska A, Tashiro A, Yuste R 2000b. Regulation of spine calcium dynamics by rapid spine motility. J Neurosci 20: 8262–8268 [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Mank M, Reiff DF, Heim N, Friedrich MW, Borst A, Griesbeck O 2006. A FRET-based calcium biosensor with fast signal kinetics and high fluorescence change. Biophys J 90: 1790–1796 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Maravall M, Mainen ZF, Sabatini BL, Svoboda K 2000. Estimating intracellular calcium concentrations and buffering without wavelength ratioing. Biophys J 78: 2655–2667 [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Markram H, Helm PJ, Sakmann B 1995. Dendritic calcium transients evoked by single back-propagating action potentials in rat neocortical pyramidal neurons. J Physiol (Lond) 485: 1–20 [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Massey PV, Johnson BE, Moult PR, Auberson YP, Brown MW, Molnar E, Collingridge GL, Bashir ZI 2004. Differential roles of NR2A and NR2B-containing NMDA receptors in cortical long-term potentiation and long-term depression. J Neurosci 24: 7821–7828 [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Matsuzaki M, Ellis-Davies GC, Nemoto T, Miyashita Y, Iino M, Kasai H 2001. Dendritic spine geometry is critical for AMPA receptor expression in hippocampal CA1 pyramidal neurons. Nat Neurosci 4: 1086–1092 [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Minelli A, Castaldo P, Gobbi P, Salucci S, Magi S, Amoroso S 2007. Cellular and subcellular localization of Na+–Ca2+ exchanger protein isoforms, NCX1, NCX2, and NCX3 in cerebral cortex and hippocampus of adult rat. Cell Calcium 41: 221–234 [DOI] [PubMed] [Google Scholar]
  78. Minta A, Kao JP, Tsien RY 1989. Fluorescent indicators for cytosolic calcium based on rhodamine and fluorescein chromophores. J Biol Chem 264: 8171–8178 [PubMed] [Google Scholar]
  79. Miyakawa H, Ross WN, Jaffe D, Callaway JC, Lasser-Ross N, Lisman JE, Johnston D 1992. Synaptically activated increases in Ca2+ concentration in hippocampal CA1 pyramidal cells are primarily due to voltage-gated Ca2+ channels. Neuron 9: 1163–1173 [DOI] [PubMed] [Google Scholar]
  80. Miyata M, Finch EA, Khiroug L, Hashimoto K, Hayasaka S, Oda SI, Inouye M, Takagishi Y, Augustine GJ, Kano M 2000. Local calcium release in dendritic spines required for long-term synaptic depression. Neuron 28: 233–244 [DOI] [PubMed] [Google Scholar]
  81. Miyawaki A, Llopis J, Heim R, McCaffery JM, Adams JA, Ikura M, Tsien RY 1997. Fluorescent indicators for Ca2+ based on green fluorescent proteins and calmodulin. Nature 388: 882–887 [DOI] [PubMed] [Google Scholar]
  82. Monyer H, Burnashev N, Laurie DJ, Sakmann B, Seeburg PH 1994. Developmental and regional expression in the rat brain and functional properties of four NMDA receptors. Neuron 12: 529–540 [DOI] [PubMed] [Google Scholar]
  83. Müller W, Connor JA 1991. Dendritic spines as individual neuronal compartments for synaptic Ca2+ responses. Nature 354: 73–76 [DOI] [PubMed] [Google Scholar]
  84. Muller M, Felmy F, Schwaller B, Schneggenburger R 2007. Parvalbumin is a mobile presynaptic Ca2+ buffer in the calyx of held that accelerates the decay of Ca2+ and short-term facilitation. J Neurosci 27: 2261–2271 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Nagai T, Sawano A, Park ES, Miyawaki A 2001. Circularly permuted green fluorescent proteins engineered to sense Ca2+. Proc Natl Acad Sci 98: 3197–3202 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Nakagawa T 2010. The biochemistry, ultrastructure, and subunit assembly mechanism of AMPA receptors. Mol Neurobiol 42: 161–184 [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Nakanishi N, Axel R, Shneider NA 1992. Alternative splicing generates functionally distinct N-methyl-d-aspartate receptors. Proc Natl Acad Sci 89: 8552–8556 [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Neher E 1998. Usefulness and limitations of linear approximations to the understanding of Ca++ signals. Cell Calcium 24: 345–357 [DOI] [PubMed] [Google Scholar]
  89. Neher E, Augustine GJ 1992. Calcium gradients and buffers in bovine chromaffin cells. J Physiol 450: 273–301 [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Nevian T, Sakmann B 2004. Single spine Ca2+ signals evoked by coincident EPSPs and backpropagating action potentials in spiny stellate cells of layer 4 in the juvenile rat somatosensory barrel cortex. J Neurosci 24: 1689–1699 [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Nevian T, Sakmann B 2006. Spine Ca2+ signaling in spike-timing-dependent plasticity. J Neurosci 26: 11001–11013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Newcomb R, Szoke B, Palma A, Wang G, Chen X, Hopkins W, Cong R, Miller J, Urge L, Tarczy-Hornoch K, et al. 1998. Selective peptide antagonist of the class E calcium channel from the venom of the tarantula Hysterocrates gigas. Biochemistry 37: 15353–15362 [DOI] [PubMed] [Google Scholar]
  93. Ngo-Anh TJ, Bloodgood BL, Lin M, Sabatini BL, Maylie J, Adelman JP 2005. SK channels and NMDA receptors form a Ca2+-mediated feedback loop in dendritic spines. Nat Neurosci 8: 642–649 [DOI] [PubMed] [Google Scholar]
  94. Nimchinsky EA, Yasuda R, Oertner TG, Svoboda K 2004. The number of glutamate receptors opened by synaptic stimulation in single hippocampal spines. J Neurosci 24: 2054–2064 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Noguchi J, Matsuzaki M, Ellis-Davies GC, Kasai H 2005. Spine-neck geometry determines NMDA receptor-dependent Ca2+ signaling in dendrites. Neuron 46: 609–622 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Oertner TG, Sabatini BL, Nimchinsky EA, Svoboda K 2002. Facilitation at single synapses probed with optical quantal analysis. Nat Neurosci 5: 657–664 [DOI] [PubMed] [Google Scholar]
  97. Plant K, Pelkey KA, Bortolotto ZA, Morita D, Terashima A, McBain CJ, Collingridge GL, Isaac JT 2006. Transient incorporation of native GluR2-lacking AMPA receptors during hippocampal long-term potentiation. Nat Neurosci 9: 602–604 [DOI] [PubMed] [Google Scholar]
  98. Regehr WG, Tank DW 1992. Calcium concentration dynamics produced by synaptic activation of CA1 hippocampal pyramidal cells. J Neurosci 12: 4202–4223 [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Reid CA, Fabian-Fine R, Fine A 2001. Postsynaptic calcium transients evoked by activation of individual hippocampal mossy fiber synapses. J Neurosci 21: 2206–2214 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Sabatini BL, Svoboda K 2000. Analysis of calcium channels in single spines using optical fluctuation analysis. Nature 408: 589–593 [DOI] [PubMed] [Google Scholar]
  101. Sabatini BL, Maravall M, Svoboda K 2001. Ca2+ signaling in dendritic spines. Curr Opin Neurobiol 11: 349–356 [DOI] [PubMed] [Google Scholar]
  102. Sabatini BL, Oertner TG, Svoboda K 2002. The life cycle of Ca2+ ions in dendritic spines. Neuron 33: 439–452 [DOI] [PubMed] [Google Scholar]
  103. Scheuss V, Yasuda R, Sobczyk A, Svoboda K 2006. Nonlinear [Ca2+] signaling in dendrites and spines caused by activity-dependent depression of Ca2+ extrusion. J Neurosci 26: 8183–8194 [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Schiller J, Helmchen F, Sakmann B 1995. Spatial profile of dendritic calcium transients evoked by action potentials in rat neocortical pyramidal neurones. J Physiol (Lond) 487: 583–600 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Schiller J, Schiller Y, Clapham DE 1998. NMDA receptors amplify calcium influx into dendritic spines during associative pre- and postsynaptic activation. Nat Neurosci 1: 114–118 [DOI] [PubMed] [Google Scholar]
  106. Schmidt H, Arendt O, Brown EB, Schwaller B, Eilers J 2007. Parvalbumin is freely mobile in axons, somata and nuclei of cerebellar Purkinje neurones. J Neurochem 100: 727–735 [DOI] [PubMed] [Google Scholar]
  107. Schneggenburger R, Zhou Z, Konnerth A, Neher E 1993. Fractional contribution of calcium to the cation current through glutamate receptor channels. Neuron 11: 133–143 [DOI] [PubMed] [Google Scholar]
  108. Seress L, Gulyás AI, Ferrer I, Tunon T, Soriano E, Freund TF 1993. Distribution, morphological features, and synaptic connections of parvalbumin- and calbindin D28k-immunoreactive neurons in the human hippocampal formation. J Comp Neurol 337: 208–230 [DOI] [PubMed] [Google Scholar]
  109. Shen W, Flajolet M, Greengard P, Surmeier DJ 2008. Dichotomous dopaminergic control of striatal synaptic plasticity. Science 321: 848–851 [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Sheng M, Cummings J, Roldan LA, Jan YN, Jan LY 1994. Changing subunit composition of heteromeric NMDA receptors during development of rat cortex. Nature 368: 144–147 [DOI] [PubMed] [Google Scholar]
  111. Simons SB, Escobedo Y, Yasuda R, Dudek SM 2009. Regional differences in hippocampal calcium handling provide a cellular mechanism for limiting plasticity. Proc Nat Acad Sci 106: 14080–14084 [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Skeberdis VA, Chevaleyre V, Lau CG, Goldberg JH, Pettit DL, Suadicani SO, Lin Y, Bennett MV, Yuste R, Castillo PE, et al. 2006. Protein kinase A regulates calcium permeability of NMDA receptors. Nature Neurosci 9: 501–510 [DOI] [PubMed] [Google Scholar]
  113. *.Smart TG, Paoletti P 2012. Synaptic neurotransmitter-gated receptors. Cold Spring Harb Perspect Biol 10.1101/cshperspect.a009662 [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Sobczyk A, Svoboda K 2007. Activity-dependent plasticity of the NMDA-receptor fractional Ca2+ current. Neuron 53: 17–24 [DOI] [PubMed] [Google Scholar]
  115. Sobczyk A, Scheuss V, Svoboda K 2005. NMDA receptor subunit-dependent [Ca2+] signaling in individual hippocampal dendritic spines. J Neurosci 25: 6037–6046 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Soler-Llavina GJ, Sabatini BL 2006. Synapse-specific plasticity and compartmentalized signaling in cerebellar stellate cells. Nat Neurosci 9: 798–806 [DOI] [PubMed] [Google Scholar]
  117. *.Südhof TC, Rizo J 2011. Synaptic vesicle exocytosis. Cold Spring Harb Perspect Biol 10.1101/a005637.cshperspect [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Sugihara H, Moriyoshi K, Ishii T, Masu M, Nakanishi S 1992. Structures and properties of seven isoforms of the NMDA receptor generated by alternative splicing. Biochem Biophys Res Commun 185: 826–832 [DOI] [PubMed] [Google Scholar]
  119. Svoboda K, Denk W, Knox WH, Tsuda S 1996. Two-photon-excitation scanning microscopy of living neurons with a saturable Bragg reflector mode-locked diode-pumped Cr:LiSrAlFl laser. Opt Lett 21: 1411–1413 [DOI] [PubMed] [Google Scholar]
  120. Svoboda K, Denk W, Kleinfeld D, Tank DW 1997. In vivo dendritic calcium dynamics in neocortical pyramidal neurons. Nature 385: 161–165 [DOI] [PubMed] [Google Scholar]
  121. Takechi H, Eilers J, Konnerth A 1998. A new class of synaptic response involving calcium release in dendritic spines. Nature 396: 757–760 [DOI] [PubMed] [Google Scholar]
  122. Tank DW, Regehr WG, Delaney KR 1995. A quantitative analysis of presynaptic calcium dynamics that contribute to short-term enhancement. J Neurosci 15: 7940–7952 [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Thiagarajan TC, Lindskog M, Tsien RW 2005. Adaptation to synaptic inactivity in hippocampal neurons. Neuron 47: 725–737 [DOI] [PubMed] [Google Scholar]
  124. Tian L, Hires SA, Mao T, Huber D, Chiappe ME, Chalasani SH, Petreanu L, Akerboom J, McKinney SA, Schreiter ER, et al. 2009. Imaging neural activity in worms, flies and mice with improved GCaMP calcium indicators. Nat Methods 6: 875–881 [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Tsien RY 1980. New calcium indicators and buffers with high selectivity against magnesium and protons: Design, synthesis, and properties of prototype structures. Biochemistry 19: 2396–2404 [DOI] [PubMed] [Google Scholar]
  126. Vicini S 1998. Functional and pharmacological differences between recombinant N-methyl-d-aspartate receptors. J Neurophysiol 79: 555–566 [DOI] [PubMed] [Google Scholar]
  127. Wang SS, Denk W, Häusser M 2000. Coincidence detection in single dendritic spines mediated by calcium release. Nat Neurosci 3: 1266–1273 [DOI] [PubMed] [Google Scholar]
  128. Watanabe S, Hong M, Lasser-Ross N, Ross WN 2006. Modulation of calcium wave propagation in the dendrites and to the soma of rat hippocampal pyramidal neurons. J Physiol (Lond) 3: 455–468 [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Waters J, Larkum M, Sakmann B, Helmchen F 2003. Supralinear Ca2+ influx into dendritic tufts of layer 2/3 neocortical pyramidal neurons in vitro and in vivo. J Neurosci 23: 8558–8567 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Westphal RS, Tavalin SJ, Lin JW, Alto NM, Fraser ID, Langeberg LK, Sheng M, Scott JD 1999. Regulation of NMDA receptors by an associated phosphatase-kinase signaling complex. Science 285: 93–96 [DOI] [PubMed] [Google Scholar]
  131. Wokosin DL, Loughrey CM, Smith GL 2004. Characterization of a range of fura dyes with two-photon excitation. Biophysical J 86: 1726–1738 [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Yasuda R, Sabatini BL, Svoboda K 2003. Plasticity of calcium channels in dendritic spines. Nat Neurosci 6: 948–955 [DOI] [PubMed] [Google Scholar]
  133. Yasuda R, Nimchinsky EA, Scheuss V, Pologruto TA, Oertner TG, Sabatini BL, Svoboda K 2004. Imaging calcium concentration dynamics in small neuronal compartments. Sci STKE 2004: p15. [DOI] [PubMed] [Google Scholar]
  134. Yuste R, Denk W 1995. Dendritic spines as basic functional units of neuronal integration. Nature 375: 682–684 [DOI] [PubMed] [Google Scholar]
  135. Yuste R, Majewska A, Cash SS, Denk W 1999. Mechanisms of calcium influx into hippocampal spines: Heterogeneity among spines, coincidence detection by NMDA receptors, and optical quantal analysis. J Neurosci 19: 1976–1987 [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Zhou Z, Neher E 1993. Mobile and immobile calcium buffers in bovine adrenal chromaffin cells. J Physiol 469: 245–273 [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Cold Spring Harbor Perspectives in Biology are provided here courtesy of Cold Spring Harbor Laboratory Press

RESOURCES